Property and Reactivity of Fluoro(silyl)acetylenes and Fluoro(stannyl

Fluoro(silyl)acetylenes and fluoro(stannyl)acetylenes underwent a radical addition reaction of THF to furnish the corresponding fluorinated cyclic eth...
0 downloads 0 Views 831KB Size
pubs.acs.org/joc

Property and Reactivity of Fluoro(silyl)acetylenes and Fluoro(stannyl)acetylenes Masahiro Shiosaki, Masashi Unno, and Takeshi Hanamoto* Department of Chemistry and Applied Chemistry, Saga University, Honjyo-machi 1, Saga 840-8502, Japan *[email protected] Received September 20, 2010

Fluoro(silyl)acetylenes and fluoro(stannyl)acetylenes underwent a radical addition reaction of THF to furnish the corresponding fluorinated cyclic ethers in moderate to good yields. These intriguing addition reaction proved to proceed via a radical reaction mechanism.

Fluorinated organic molecules would be one of the most important classes of halogenated compounds with rare occurrence in nature.1 They often play a significant role in the modern drug discovery process, which should make the development of new synthetic methodologies to facilitate their preparation worthwhile. 2 Therefore, a number of methods have been reported to synthesize a variety of fluoroalkanes, fluoroalkenes, and fluoroarenes due to their useful applications. In contrast to such main stream, study on fluoroalkyne chemistry has lagged far behind.3 One of the reasons may be attributed to the instability and explosive nature of fluoroacetylenes in previous reports.4 We have recently undertaken studies aimed at such a fluoroacetylene (1) (a) Uneyama, K. In Organofluorine Chemistry; Blackwell: Oxford, UK, 2006. (b) Chambers, R. D. In Fluorine in Organic Chemistry; Blackwell: Oxford, UK, 2004. (c) Hiyama, T. In Organofluorine Compounds; Springer: New York, 2000. (d) Hudlicky, M.; Pavlath, A. E. In Chemistry of Organic Fluorine Compounds II, ACS Monograph 187; American Chemical Society: Washington, DC, 1995. (2) (a) Kirk, K. L. Org. Process Res. Dev. 2008, 12, 305. (b) Purser, S.; Moore, P. R.; Swallow, S.; Gouverneur, V. Chem. Soc. Rev. 2008, 37, 320. (c) M€ uller, K.; Faeh, C.; Diederich, F. Science 2007, 317, 1881. (d) Isanbor, C.; O’Hagan, D. J. Fluorine Chem. 2006, 127, 303. (3) Hopf, H.; Witulski, B. In Modern Acetylene Chemistry; Stang, P. J., Diederich, F., Eds.; VCH: New York, 1995; Chapter 2. (4) (a) B€ urger, H.; Sommer, S. J. Chem. Soc., Chem. Commun. 1991, 456. (b) Runge, A.; Sander, W. W. Tetrahedron Lett. 1990, 31, 5453. (c) Chambers, R. D.; Shepherd, T.; Tamura, M.; Bryce, M. R. J. Chem. Soc., Chem. Commun. 1989, 1657. (d) Kende, A. S.; Fludzinski, P. J. Org. Chem. 1983, 48, 1384. (e) Viehe, H. G.; Delavarenne, S. Y. Chem. Ber. 1970, 103, 1216. (f ) Delavarenne, S. Y.; Viehe, H. G. Chem. Ber. 1970, 103, 1209. (g) Delavarenne, S. Y.; Viehe, H. G. Chem. Ber. 1970, 103, 1198. (h) Viehe, H. G.; Merenyi, R.; Oth, J. F. M.; Sender, J. R.; Valange, P. Angew. Chem., Int. Ed. 1964, 3, 755. (i) Viehe, H. G.; Merenyi, R.; Oth, J. M. F.; Valange, P. Angew. Chem., Int. Ed. 1964, 3, 746. ( j) Middleton, W. J.; Sharkey, W. H. J. Am. Chem. Soc. 1959, 81, 803.

8326

J. Org. Chem. 2010, 75, 8326–8329

family and reported a synthetic method for the facile preparation of fluoro(triisopropylsilyl)acetylene 1a.5 As we noticed that 1a turned out to be stable only in a hexane solution at low temperature, its further synthetic applications remained unexplored. After we started to investigate the properties and reactivity of fluoro(silyl)acetylenes, we successfully isolated fluoro(triisopropylsilyl)acetylene 1a in the pure form and measured its 13C NMR for the first time. In addition, we serendipitously encountered the addition reaction of THF to other analogous fluoro(silyl)acetylenes and fluoro(stannyl)acetylenes except for 1a. In this paper, we disclose an intriguing behavior of fluoro(silyl)acetylenes and fluoro(stannyl)acetylenes and propose the possible reaction mechanism of the fluorinated cyclic ethers. Although we have previously established the facile preparation of 1a from 1,1-difluoroethylene, the fluoroacetylene was only stable in a hexane solution. Thus, concentration of the hexane solution containing 1a resulted in the decomposition of 1a. However, the different perspective for this outcome provided us a hint that hexane as a solvent might be responsible for stabilization of 1a. On the basis of this idea we carried out purification of 1a through distillation from the corresponding hexadecane solution containing 1a (we confirmed that hexadecane has a higher boiling point in comparison with that of 1a by GC-MS in advance) using a short path distillation apparatus. We determined that the desired acetylene 1a was successfully distilled at 30-35 °C (bath temperature) at 40 Pa as a colorless oil. To our delight, the isolated compound 1a was stable enough to execute the following measurement. We immediately attempted to measure the 13C NMR spectrum of 1a. Although 1a in CDCl3 started slowly to decompose at -10 °C, over ten accumulation times we were able to perform the 13C NMR measurement without any detectable decomposition of 1a. The 13C NMR spectrum of 1a is shown in Figure 1.

Interestingly, one acetylene carbon signal appeared at 19.09 ppm (d, J = 17.4 Hz) and the other signal appeared at 106.91 ppm (d, J = 337.6 Hz). On the basis of the coupling constant of carbon atom-fluorine atom, the former was assigned to the sp-carbon attached to silicon and the latter was assigned to the sp-carbon attached to fluorine. Although we did not have any rational reason to explain these two sp-carbon chemical shifts at the present stage, this is the first 13 C NMR spectrum observation of sp-carbon attached directly to fluorine. In addition, the elemental analysis of 1a was successfully conducted to give satisfactory results (elemental analysis calcd for C11H21FSi: C 65.94, H 10.56; found: C 66.19, H 10.82) for the first time, although exposure of the colorless oil 1a to air resulted in coloration to the pale yellow oil. To assist the interpretation of the unexpected chemical shift values shown above, we have conducted quantum (5) Hanamoto, T.; Koga, Y.; Kawanami, T.; Furuno, H.; Inanaga, J. Angew. Chem., Int. Ed. 2004, 43, 3582.

Published on Web 11/10/2010

DOI: 10.1021/jo1018602 r 2010 American Chemical Society

JOC Note

Shiosaki et al. SCHEME 1.

FIGURE 1.

13

TABLE 1.

13

exptl HF MP2 B3LYP

C NMR of 1a.

C NMR Chemical Shifts (ppm) for sp-Carbon Atoms δC1 δC2 106.9 115.8 111.8 119.1

19.1 17.0 17.2 16.4

chemical calculations for 1a using Gaussian03.6 Geometry optimizations were performed at the B3LYP/6-31G** level, and the optimized structures were used for the subsequent calculations of molecular properties. The GIAO nuclear magnetic shielding values were calculated at all atomic positions at the HF, MP2, and B3LYP levels of theory using the 6-311þG(2d,p) basis set. The chemical shifts for the carbon atoms were obtained by subtracting their chemical shielding values from the one calculated for TMS, which is 192.6, 192.9, or 182.8 ppm for HF, MP2, or B3LYP, respectively. Table 1 summarizes the observed and calculated chemical shifts for the sp-carbon atoms. As shown in the table, the present quantum chemical calculations reproduce the observed chemical shifts. We turned our attention to substrate scope and chose fluoro(dimethylphenylsilyl)acetylene 1b as an analogous molecule of 1a. We attempted to prepare 1b in the similar manner. After complete addition of the components at -78 °C, the reaction mixture was allowed to stand at room temperature overnight under exclusion of light.7 At this time we encountered the following unexpected finding. When we checked the reaction progress by GC-MS analysis prior to quench, the original peak corresponding to 1b [m/z 163 (Mþ - Me), M = 178 (C10H11FSi)] almost disappeared and (6) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, € Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, O.; Revision A.2; Gaussian, Inc., Wallingford CT, 2009. (7) The same reaction in the light caused formation of unidentified byproduct and decreased the yield of 2b.

Synthesis of 2b and 3b

a new major peak emerged at m/z 234 along with minor peaks. After the usual workup, the residual oil was separated by silica gel column chromatography to afford the corresponding products. The main product showed one vinyl proton signal at 5.07 ppm (δ, J = 34.1 Hz) in the 1H NMR spectrum, one vinyl fluorine signal at -85.9 ppm (dd, J = 33.9, 24.2 Hz) in the 19F NMR spectrum, and two vinyl carbon signals at 99.7 (d, J = 3.1 Hz) and 167.69 ppm (d, J = 279.6 Hz) in the 13C NMR spectrum, respectively. On the basis of these NMR data, this molecule was assigned to be (E)-[2-fluoro-2-(tetrahydrofuran-20 -yl)vinyl]dimethylphenylsilane 2b.8 On the other hand, the minor (Z)-isomer 3b showed one vinyl proton signal at 4.96 ppm (d, J = 61.7 Hz) in the 1H NMR spectrum and one vinyl fluorine signal at -95.0 ppm (dd, J = 62.0, 8.9 Hz) in the 19F NMR spectrum. The characteristic H-F coupling constant of each stereoisomer justified the structural assignment. The E/Z ratio of the stereoisomers (2b/3b) was determined to be 86:14 by GC-MS analysis, using a crude reaction mixture prior to chromatographic purification (Scheme 1). These unexpected results prompted us to examine substrate scope. Various chlorosilanes and chlorostannanes were subjected to investigation. Although the addition reactions proceeded giving the corresponding adducts 2c-h, the yields were generally modest under similar conditions [Table 2, Method A (left column)]. Interestingly, sterically demanding chlorotriphenylsilane scarcely underwent the first substitution reaction of lithium fluoroacetylide with the result that a trace amount of the corresponding adduct was detected by GC-MS (entry 6). These unsatisfactory results led us to reexamine the reaction conditions for the preparation of 2b. After many attempts,9 we eventually found that the key factor contributing to higher yield probably depended on the reaction concentration. Thus, after the complete generation of 1b at -78 °C was confirmed, the reaction mixture was gradually warmed to -40 °C. At this temperature, an excessive amount of THF was carefully added to the mixture. This diluted reaction mixture was gradually warmed to room temperature while being stirred, which resulted in acceptable yield (up to 78%). Once the optimum reaction conditions had been identified, the remaining substrates were again subjected to examination. All of the reactions proceeded well with moderate to high yields [Table 2, Method B (right column)]. Although the yields were improved in this way in almost every case, the problem of stereoselectivity still remains unsolved. To gain the reaction progress in detail we monitored the reaction of 1b with the reaction temperature by GC-MS analysis. The first appearance of the peak 2b (8) Muramatsu, H.; Inukai, K.; Iwata, Y.; Murakami, S. Bull. Chem. Soc. Jpn. 1967, 40, 1284. (9) Use of THF as a solvent is essential to successful preparation of fluoro(silyl)acetylene.

J. Org. Chem. Vol. 75, No. 23, 2010

8327

JOC Note TABLE 2.

Reaction of Fluoro(silyl)acetylene and Fluoro(stannyl)acetylene

a Isolated yield. bDetermined by GC-MS analysis in the crude reaction mixture otherwise noted. cUnder the preliminary reaction conditions. d Reaction concentration (0.17 M). eUnder the optimized reaction conditions. fReaction concentration (0.10 M). gReaction concentration (0.07 M). hDetermined by NMR after column chromatography. iThe corresponding fluoro(triphenylsilyl)acetylene was scarcely detected by GC-MS analysis.

along with 1b was observed at around -20 °C. Thus, 1b remained intact in THF at least below ca. -30 °C, and was subjected to the addition reaction of THF at ca. -20 °C, giving 2b and 2c. Our next task is to clarify whether this reaction proceeded via an ionic or a radical process. We conducted the following two reactions under different conditions. Thus, after preparing 1b at -78 °C again, first we slowly added an excessive amount of THF (5 mL) and water (10 equiv to 1b) to this solution at -40 °C. The mixture was gradually warmed to room temperature after the complete addition. In this case, the addition reaction smoothly proceeded to afford the corresponding products 2b and 3b in essentially the same manner [82%, 2b:3b = 81:19, compare to entry 1 (Method B) in Table 1 (78%, 2b:3b = 82:18)]. On the other hand, after preparing 1b at -78 °C, the addition of 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO, 1.1 equiv to 1b) instead of water at the same time completely retarded the addition reaction to afford no THF adducts at -20 °C. Moreover, even at room temperature, the peak of 1b was observed essentially intact by GC-MS in the reaction mixture for a short time. TEMPO gradually decomposed at this temperature under these conditions, which may cause decomposition of 1b, finally giving the complex reaction mixtures. Surprisingly, even at the final 8328

J. Org. Chem. Vol. 75, No. 23, 2010

Shiosaki et al. SCHEME 2.

Plausible Reaction Mechanism

stage no formation of 2b and 3b was detected in the complex mixtures by GC-MS. Furthermore, EPR studies of the reaction were briefly attempted to confirm its radical mechanism. We monitored the EPR spectra of the reaction solution containing 1b and 5,5-dimethyl-1-pyroline N-oxide (DMPO) as a radical trap reagent with the reaction temperature. We observed the multiplet line with a g value of 2.0066, which corresponded to 5,5-dimethyl-2-substituted-1-pyrrolidine N-oxide radical. These peaks corresponded to the triplet hyperfine structure by an N nuclei (l = 1) and the doublet hyperfine structure by one adjacent H nuclei (l = 1/2). The former hyperfine coupling constant (AN) was 1.38 mT and the latter one (AH) was 1.72 mT. However, no further information about the substituent at the 2-position on the pyrrolidine ring was obtained under our experimental conditions. On the basis of these results and previous findings in another group,10 our proposed radical mechanism is shown in Scheme 2. Initially, a trace amount of inevitable and/or adventitious oxygen generated the corresponding THF radical 4. The radical 4 almost regioselectively attacked at the β-carbon of 1b to yield the corresponding fluorovinyl radical 5. The radical 5 abstracted the R-hydrogen of THF to afford the products (2b and 3b) and to regenerate the THF radical 4. As the result, the complete radical cycle mechanism should be successfully accomplished. In summary, we have demonstrated fluoro(triisopropylsilyl)acetylene 1a and the first successful 13C NMR measurement of the radical reaction of fluoro(silyl)acetylenes (1b-f) and fluoro(stannyl)acetylenes (1h and 1i) with THF. We also have verified the stability of these fluoro(silyl)acetylenes and fluoro(stannyl)acetylenes in THF in the absence of oxygen. Although their final decomposition mechanism is not clear at the present stage, these findings would contribute a new aspect of fluoroacetylene chemistry. Experimental Section [2-Fluoro-2-(tetrahydrofuran-20 -yl)vinyl]dimethylphenylsilane (2b and 3b). A 25-mL two-necked flask equipped with a magnetic stir bar, a stopcock, and a three-way stopcock was charged with 2.5 mL of THF under argon. To this solution was added sBuLi (1.04 M in cyclohexane-hexane solution, 1.4 mL, 1.46 mmol, 2.2 equiv) dropwise via syringe at -78 °C. After being stirred for (10) Bergstrom, D. E.; Ng, M. W.; Wong, J. J. Chem. Soc., Perkin Trans. 1 1983, 741.

Shiosaki et al. an additional 5 min, the solution was cooled to -120 °C via liquid N2/ethanol bath. At this temperature, argon was replaced with 1,1-difluoroethylene (balloon). The mixture was gradually warmed to -85 °C. HMPA (6.0 μL, 5 mol %) at -85 °C and chlorodimethylphenylsilane (110 μL, 0.67 mmol) at -78 °C were successively added to the solution via syringe. After the mixture was stirred for 10 min, the reaction flask was covered with aluminum foil. The mixture was slowly warmed to -40 °C while an excess amount of 1,1-difluoroethylene was gradually let out through the drying glass tube filled with calcium chloride. When the reaction mixture reached -40 °C, the additional THF (5.0 mL) was carefully added to the solution. After removal of the cooling bath at -30 °C, the reaction mixture was allowed to stand overnight. To the resulting solution was carefully added hexane and water, successively. After the resultant mixture was extracted with hexane, additional extraction was repeated twice. The combined solution was dried over sodium sulfate. The solution was concentrated in vacuo and the residual oil was quickly purified by chromatography on a short silica gel column covered with aluminum foil (hexane:ethyl acetate:triethylamine = 200:4:1 as an eluent) to give a mixture of 2b and 3b [78% (130.7 mg), E/Z = 82:18]. The pure sample 2b was obtained even at the expense of the yield from partial fractions by the second chromatography on longer silica gel column of the mixture of 2b and 3b under the same separation conditions. 2b: Rf 0.49; hexane:ethyl acetate:triethylamine = 200:4:1; elemental analysis calcd for C14H19FOSi: C 67.16, H 7.65; found: C 67.19, H 7.69; δH = 0.40 (3H, s), 0.42 (3H, s), 1.70-2.00 (4H, m),

JOC Note 3.65-3.90 (2H, m), 4.43 (1H, dt, JHF = 25.0 Hz, JHH = 7.3 Hz), 5.07 (1H, d, JHF = 34.1 Hz), 7.30-7.50 (5H, m) ppm; δF = -85.9 (dd, JFH = 34.1, 25.0 Hz) ppm; δC = -0.86 (d, JCF = 1.9 Hz), -0.69 (d, JCF = 1.9 Hz), 26.3 (s), 29.2 (s), 69.2 (s), 76.1 (d, JCF = 32.4 Hz), 99.7 (d, JCF = 3.1 Hz), 127.9 (s), 129.2 (s), 133.6 (s), 138.5 (d, JCF = 1.9 Hz), 167.7 (d, JCF = 279.6 Hz) ppm; ν = 3069, 2956, 2873, 1652, 1429, 1112, 1059, 791, 700 cm-1; GC-MS (m/z) 235 [Mþ - 15 (Me), 47], 193 (46), 155 (8), 129 (76), 105 (23), 77 (100), 53 (16). 3b: Rf 0.30; hexane: ethyl acetate:triethylamine = 200:4:1; δH = 0.40 (3H, s), 0.41 (3H, s), 1.85-2.16 (4H, m), 3.80-4.00 (2H, m), 4.39 (1H, dt, JHF = 8.3 Hz, JHH = 4.8 Hz), 4.96 (1H, d, JHF = 61.7 Hz), 7.30-7.50 (5H, m) ppm; δF = -95.0 (dd, JFH = 61.7, 8.3 Hz) ppm; ν = 3071, 2957, 1662, 1428, 1249, 1114, 1078, 1060, 834, 731, 699 cm-1; GC-MS (m/z) 235 [Mþ - 15 (Me), 0.9], 173 (4), 139 (100), 129 (60), 91 (65), 79 (42), 77 (84).

Acknowledgment. We greatly thank both Tosoh F-Tech, Inc. for a gift of 1,1-difluoroethylene and Chisso Corp. for a gift of chlorolsilanes. We thank Prof. Junji Inanaga, Prof. Hiroshi Furuno, and Prof. Satoru Karasawa at Kyushu University and Prof. Masayuki Koikawa at Saga University for their valuable discussions and experimental support. Supporting Information Available: Characterization of new compounds 2c-i, and NMR spectral for all new compounds. This material is available free of charge via the Internet at http:// pubs.acs.org.

J. Org. Chem. Vol. 75, No. 23, 2010

8329